characterizing_the_distribution

advertisement
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
Characterizing the distribution and rates of microbial sulfate reduction at Middle Valley
hydrothermal vents
Kiana L. Frank1, Daniel R. Rogers1, Heather C. Olins1, Charles Vidoudez1 and Peter R. Girguis1*
1 = Department of Organismic and Evolutionary Biology, Harvard University, 16 Divinity Avenue,
Cambridge, MA 02138-2020.
* = send correspondence to pgirguis@oeb.harvard.edu
Running Title: Sulfate reduction rates at hydrothermal vents
30
31
1
32
ABSTRACT
33
Few studies have directly measured sulfate reduction at hydrothermal vents, and
34
relatively little is known about how environmental or ecological factors influence rates of
35
sulfate reduction in vent environments. A better understanding of microbially mediated sulfate
36
reduction in hydrothermal vent ecosystems may be achieved by integrating ecological and
37
geochemical data with metabolic rate measurements. Here we present rates of microbially
38
mediated sulfate reduction from three distinct hydrothermal vents in the Middle Valley vent
39
field along the Juan de Fuca Ridge, as well as assessments of bacterial and archaeal diversity,
40
estimates of total biomass and the abundance of functional genes related to sulfate reduction,
41
and in situ geochemistry. Maximum rates of sulfate reduction occurred at 90°C in all three
42
deposits. Pyrosequencing and functional gene abundance data reveal differences in both
43
biomass and community composition among sites, including differences in the abundance of
44
known sulfate reducing bacteria. The abundance of sequences for Thermodesulfovibro-like
45
organisms and higher sulfate reduction rates at elevated temperatures, suggests that
46
Thermodesulfovibro-like organisms may play a role in sulfate reduction in warmer
47
environments. The rates of sulfate reduction presented here suggest that - within anaerobic
48
niches of hydrothermal deposits - heterotrophic sulfate reduction may be quite common and
49
can contribute to secondary productivity, underscoring the potential role of this process in both
50
sulfur and carbon cycling at vents.
51
Keywords: Hydrothermal Vent/Microbial Ecology/Primary productivity/Sulfate Reduction
52
Subject Category: Geomicrobiology and microbial contributions to geochemical cycles
53
2
54
INTRODUCTION
55
Deep sea hydrothermal vent ecosystems are complex dynamic habitats characterized by
56
steep gradients in temperature and geochemistry (Jannasch & Mottl, 1985). In these habitats,
57
as hot hydrothermal fluid mixes with cold seawater, the precipitation of minerals creates large
58
and complex hydrothermal chimney deposits. Within these permeable mineral structures, the
59
continued mixing of chemically reduced, vent-derived fluids and oxidized seawater provides
60
favorable conditions that support the growth of endolithic microbial communities (Schrenk et
61
al. 2003)
62
Sulfide oxidation is considered to be one of the most important microbial
63
chemosynthetic pathways at ridge ecosystems, as evidenced by the ubiquity of sulfide oxidizing
64
Epsilon- and Gammaproteobacteria at ridge environments (Nakagawa et al. 2004; Huber et al.
65
2007; Nakagawa & Takai, 2008; Nakagawa et al. 2005; Campbell et al. 2006). To date,
66
significantly less attention has been paid to the distribution and magnitude of sulfate reduction
67
at vents, though sulfate reducing bacteria and archaea have frequently been isolated from deep
68
sea hydrothermal environments (Houghton et al. 2007; Audiffrin et al. 2003; Alazard et al.
69
2003; Jannasch et al. 1988; Blöchl et al. 1997). Moreover, analyses of functional genes that
70
express key proteins required for sulfate reduction suggest there is a high diversity of sulfate
71
reducing organisms at vents, higher than predicted via 16S rRNA gene analyses alone
72
(Nakagawa et al. 2004; Nercessian et al. 2005).
73
From a biogeochemical and bioenergetic perspective, both sulfide oxidation and sulfate
74
reduction would be favored at hydrothermal vents, though to varying degrees as a function of
75
environmental chemistry. Sulfide oxidation is most favorable when coupled to oxygen or nitrate
3
76
as an electron acceptor (Amend & Shock 2001). Around vents, sulfide is typically in µM to mM
77
concentrations (Butterfield, et al. 1994; Butterfield et al. 1994), while oxygen and nitrate are
78
around 110 and 40 µM respectively (Johnson et al. 1986). In contrast, sulfate reduction is highly
79
favored in anoxic niches at vents, as it is in other marine anaerobic environments (Muyzer &
80
Stams, 2008). Here, as in most marine systems, sulfate is abundant at 28 mM two to three
81
orders of magnitude higher than oxygen At vents, sulfate reduction would occur in regimes
82
where seawater-derived sulfate is still present but oxygen is absent, e.g. within hydrothermal
83
vent deposits. Sulfate reducing microorganisms commonly use hydrogen and/or dissolved
84
organic matter as electron donors, both of which are found within hydrothermal fluids (Lang et
85
al. 2006; Cruse & Seewald, 2006). Sulfate reduction –as a function of its extent and magnitude-
86
could readily influence the cycling of sulfur and sulfur isotopes, as well as carbon, within
87
hydrothermal environments.
88
To date, studies have quantified rates of sulfate reduction in hydrothermal-influenced
89
sediments (Weber & Jorgensen, 2002; Jorgensen et al. 1992; Elsgaard et al. 1994; Kallmeyer &
90
Boetius, 2004; Elsgaard et al. 1994; Elsgaard et al. 1995) and isolated vent microorganisms
91
(Hoek et al. 2003). In contrast to the numerous studies of sulfate reduction in marine sediments
92
(Canfield 1989), studies of sulfate reduction in hydrothermal deposits are few (Bonch-
93
Osmolovskaya et al. 2011), due in part to the challenges associated with sampling and studying
94
the heterogeneous and consolidated sulfide deposits typical of hydrothermal vent chimneys.
95
Here we present rates of microbially mediated sulfate reduction from three distinct,
96
active hydrothermal “chimneys” found in the Middle Valley field along the Juan de Fuca Ridge,
97
as well as assessments of bacterial and archaeal diversity, estimates of total biomass and the
4
98
abundance of functional genes related to sulfate reduction, and in situ geochemistry. These
99
analyses further our understanding of sulfate reduction (including rates, diversity and
100
distribution of known sulfate-reducing microbes) in vent ecosystems. Moreover, they
101
underscore the potential role of heterotrophic sulfate reduction in hydrothermal systems, and
102
constrain their potential influence on both sulfur and carbon cycling,
103
104
105
106
METHODS
Geologic Setting and Sampling of hydrothermal deposits
107
Middle Valley (48°27’N, 128°59’ W) is an intermediate spreading, axial rift valley,
108
located along the Endeavor Segment of the Juan de Fuca Ridge in the Northwest Pacific ocean.
109
Layers of continental-derived sediments characteristically cover Middle Valley, though the
110
hydrothermal vents remain prominent above the sediments. Hydrothermal deposits were
111
collected from 3 active hydrothermal spires during dive 4625 with the HOV Alvin (R/V Atlantis
112
expedition AT15-67, July 2010) and brought to the surface in a sealed, temperature-insulated
113
polyethylene box. Samples were recovered from actively venting sulfide deposits at Needles
114
(48.45778, -128.709, 2412.212 m, Tmax=123°C), Dead Dog (48.45603,-128.71, 2405.268 m,
115
Tmax=261°C), and Chowder Hill (48.455543, -128.709, 2398.257 m, Tmax =261°C) vents. Once on
116
board ship, samples were directly transferred to sterile anaerobic seawater and
117
handled/processed using appropriate sterile microbiological techniques. Subsamples were
118
immediately transferred to gastight jars (Freund Container Inc.), filled with sterile anaerobic
119
seawater containing 2 mM sodium sulfide at pH 6, and stored at 4°C. Upon return to the
120
laboratory, all samples were provided with fresh 2 mM sulfidic, anaerobic seawater every 8 to
121
12 weeks and were kept in the dark and 4°C prior to incubation.
5
122
Vent fluid volatile geochemistry via in situ mass spectrometry
123
In situ concentration of dissolved volatiles (H2S, H2, CO2, O2, and others.) were measured
124
at each site with an in situ mass spectrometer (ISMS) as previously described (Wankel et al.
125
2011). Briefly, dissolved volatiles were quantified in situ by sampling vent effluent for up to 10
126
minutes, until partial pressures reached steady state (data was monitored in real time within
127
the submersible). Concentrations were determined from empirically derived calibrations and
128
validated by comparison with discrete samples collected using titanium gastight samplers.
129
Measuring sulfate reduction rates
130
Hydrothermal deposits were homogenized in a commercial blender (Xtreme™ blender,
131
Waring Inc.) under a nitrogen atmosphere. Anaerobic homogenization was designed to
132
minimize fine-scale geochemical and microbial heterogeneity and facilitate more accurate
133
experimental replication. Hydrothermal homogenate (made up of both mineral deposit and
134
interstitial fluid)) was aliquoted volumetrically (7.5 mL, ca. 29 g wet weight and ca. 20 g dry
135
weight) into Balch tubes in an anaerobic chamber. The tubes were supplemented with 15 mL of
136
sterile artificial vent fluid media designed to mimic the geochemical conditions within a sulfide
137
deposit (pH 6, 14 mM SO42-, 2.3 mM NaHCO3, 1 mM H2S, and 10 μM each of pyruvate, citrate,
138
formate, acetate, lactate). Organic acid concentrations are comparable to those measured in
139
situ (Lang et al. 2006). Sufficient 35SO42- was added to achieve 555 kBq (15 μCi) of activity. Due
140
to technical difficulties with post processing methodology, shipboard incubations using fresh
141
material were not successful. The data presented here were generated using samples that had
142
been kept at 4 °C and refreshed with vent-like effluent (as described above) for one year.
143
Samples were incubated anaerobically for 7 days at 4, 30, 40, 50, 60, 80 and 90°C. Controls for
6
144
sulfate reduction consisted of samples amended with 28 mM molybdate, a competitive
145
inhibitor of sulfate reduction (Saleh et al. 1964; Newport & Nedwell, 1988). Six biological
146
replicates were run for each treatment, and three biological replicates for each control. Upon
147
completion, reactions were quenched with the injection of 5 mL 25% zinc acetate (which is ~20-
148
fold more Zinc than the maximum sulfide concentration), and all samples were frozen at -20° C
149
for further analysis.
150
To determine sulfate reduction rates, samples were thawed and the supernatant was
151
removed and filtered through a 0.2 μm syringe filter. The crushed deposits that remained in the
152
tube were washed three times with deionized water to remove any remaining sulfate. One
153
gram (wet weight) of crushed deposit was analyzed via chromium distillation (see Supplemental
154
Methods) and sulfate reduction rates (SRR) were calculated as in (Fossing & Jorgensen,
155
1989)using the following calculation.
156
SRR =
Eq.1
157
Where nSO42- is the quantity (in moles) of sulfate added to each incubation (14 mM * 15 mL =
158
210 μmol), a is the activity (dpm) of the trapped sulfide, 1.06 is the fractionation factor
159
between the sulfide and sulfate pools, A is the activity of the sulfate pool at the completion of
160
the incubation and t is the incubation time (days). The rates are presented in units of nmol S g -1
161
day-1.
162
DNA Extraction
163
Immediately prior to conducting the rate experiments, a subsample of homogenized
164
hydrothermal deposit was removed and frozen at -80° C for molecular analysis. DNA was
7
165
extracted from this crushed deposit sample with a protocol modified from (Santelli et al. 2008).
166
Subsamples were washed with 0.1 N HCl, followed by two rinses with a sterile solution
167
containing 10 mM Tris (pH 8.0) and 50 mM EDTA. A known mass of material was added to
168
PowerSoil beadbeating tubes(MoBio Laboratories, Carlsbad CA), incubated at 70°C for 10
169
minutes, and then amended with 200 ng of poly-A. Subsamples were subjected to beadbeating,
170
followed by three cycles of freeze-thaw steps to further lyse cells. Nucleic acids were extracted
171
using hot phenol (60°C for 3 min.), followed by two chloroform:isoamyl separations and
172
precipitated with ethanol. DNA was resuspended in TE (pH 8.0) and quantified using the Qubit™
173
fluorometer (Life Technologies, Grand Island, NY).
174
Enumeration of gene abundance via quantitative PCR
175
Quantitative PCR (qPCR) was used to determine the abundance of bacterial and archaeal
176
16S rRNA genes. In addition, qPCR was used to enumerate the abundance of sulfate reducing
177
prokaryotes by amplifying the adenosine 5‘-phosphosulfate reductase (aprA) gene with primers
178
targeting sulfate reducing bacteria and archaea (Christophersen et al. 2011). Primers specific to
179
bacterial dissimilatory sulfite reductase (dsrA) (Kondo et al. 2004) and Deltaproteobacteria 16S
180
rRNA genes (Stults et al. 2001) provide alternate estimates of sulfate reducing bacteria
181
populations. Quantification was performed in triplicate with the Stratagene MX3005p qPCR
182
System (Agilent Technologies) using the Perfecta SYBR FastMix with low ROX (20 µL reactions,
183
Quanta Biosciences, Gaitherburg, MD), specific primers and annealing temperatures (Table 1)
184
and 10 ng of template gDNA. The temperature program for all assays was 94°C for 10 minutes,
185
35 cycles of 94°C for 1 minute, the annealing temperature for 1 minute (Table 1), extension at
186
72°C for 30 seconds and fluorescence read after 10 seconds at 80°C. Following amplification,
8
187
dissociation curves were determined across a temperature range of 55°C to 95°C. Ct values for
188
each well were calculated using the manufacturer’s software. Plasmids containing bacterial and
189
archaeal 16S rRNA and functional gene inserts (amplified from Arcobacter nitrofigulis (ATCC
190
33309), Methanosarcina acidovorans and Desulfovibrio vulgaris Hildenborough (ATCC 29579/
191
NCIMB 8303/ AE017285) respectively) were used as standards for calibration (see
192
Supplemental Methods for more detail).
193
Sequencing and Phylogenetic Analysis via 454 pyrosequencing
194
DNA samples were sequenced by Research and Testing Laboratory (Lubbock, TX) using 454
195
pyrotag methods similar to those described previously (Dowd et al. 2008). All samples were
196
sequenced using a 454FLX instrument (Roche Inc.) with Titanium™ reagents. The resulting
197
bacterial and archaeal 16S rRNA and drsB genes (primers in Table 1) datasets were analyzed via
198
Mothur (Schloss et al. 2009). Sequences were trimmed, quality checked, aligned to the SILVA-
199
compatible alignment database reference alignment (dsrB gene datasets were aligned to a dsrB
200
gene database generated from the Ribosomal Database Project (RDP)), analyzed for chimeras,
201
classified against the Greengenes99 database and clustered in to OTUs (see Supplemental
202
Methods for more detail). Rarefaction curves were used to examine the number of OTUs as a
203
function of sampling depth. Alpha diversity was assessed by generating values from the Chao1
204
richness estimators and the inverse Simpson diversity index.
205
Sequence Accession numbers
206
The 16S rRNA and dsrB gene sequences reported in this study have been submitted to
207
Sequence Read Archive under the accession numbers SRX154520 through SRX154528.
208
209
RESULTS
9
210
211
Physical and geochemical characteristics of the study sites
212
The hydrothermal deposits sampled from Middle Valley were all relatively friable and
213
were composed predominantly of anhydrite (CaSO4, M. Tivey, pers. comm). Chowder Hill and
214
Dead Dog had the highest observed venting fluid temperatures (measured in situ at 261°C),
215
followed by Needles (123°C). In situ measurements of dissolved hydrogen sulfide (H2S) revealed
216
significant differences in hydrothermal fluid composition among hydrothermal deposits.
217
Unfortunately the inline pH probe with the ISMS malfunctioned during the dive. Using
218
previously reported pH values (Butterfield et al. 1994), Chowder Hill would have the highest in
219
situ measurement of total sulfide (3.9 mM), followed by Dead Dog (2.2 mM), and Needles (0.59
220
mM) (Table 2). These concentrations are within the same magnitude of previously reported H2S
221
in focused vent fluids at Middle Valley (Butterfield et al. 1994). Chowder Hill did exhibit the
222
highest in situ concentration of hydrogen (1.86 mM) followed by Dead Dog (1.66 mM) and
223
Needles (~1.42 mM). These values are also consistent with previous studies (Cruse & Seewald
224
2006), as well as gastight samples collected and analyzed shipboard (M. Lilley, pers. comm).
225
Sulfate Reduction Rates
226
Among all samples, sulfate reduction was observed at temperatures between 4°C to
227
90°C (Figure 1). Maximal rates of sulfate reduction were observed between 88-90°C (2670 nmol
228
g-1 day-1 at Needles, 1090 nmol g-1 day-1 at Chowder Hill, and 142 nmol g-1 day-1 at Dead Dog;
229
Figure 1). Notably, the highest sulfate reduction rates were observed from Needles samples,
230
which were ~20-fold higher than those observed at Dead Dog, and ~2-fold greater than at
231
Chowder Hill. Many of the rates exhibit large deviations due to the high variability among the
232
biological replicates, most likely due to persistent mineralogical and microbiological
10
233
heterogeneity across incubations, even after homogenization. Sulfate reduction was also
234
observed in molybdate amended experiments, though we suspect that molybdate was
235
scavenged by minerals that attenuated the effect of the inhibitor as has been previously
236
observed in metal-rich environments (Bostick et al. 2003; Xu et al. 2006).
237
Quantification of Taxonomic and Functional Genes
238
The abundance of total bacteria, archaea (16S rRNA genes), sulfate reducing
239
prokaryotes (aprA gene) and sulfate reducing bacteria (dsrA and Deltaproteobacteria specific
240
16S rRNA genes) were investigated in each deposit by quantitative PCR (Figure 2). Microbial
241
density (as estimated by 16S rRNA gene copies g-1 mineral) was greatest at Needles and lowest
242
at Dead Dog. Microbial communities at each site were dominated by archaea (Figure 2A), with
243
Needles showing the highest ratio of archaea to bacteria (227:1 as compared to 14:1 at Dead
244
Dog or 17.5:1 at Chowder Hill). Assuming an average of 4.19 copies of 16S rRNA gene per
245
bacterium and 1.71 copies of 16S rRNA gene per archaeon genome (Lee et al. 2009;
246
Klappenbach et al. 2001), Needles hosts a microbial community of 4.12 x 108 cells g-1 sample, 3
247
orders of magnitude higher than Chowder Hill (8.96 x 105 cells g-1 sample) and Dead Dog (5.65 x
248
105 cells g-1 sample).
249
16S rRNA gene primers specifically targeting ribotypes allied to Desulfovibrio,
250
Desulfomicrobium,
Desulfuromusa,
and
Desulfuromonas
were
used
to
enumerate
251
Deltaproteobacteria known to mediate sulfate reduction in many marine systems (Stults et al.
252
2001). However, given the difficulty in amplifying 16S rRNA genes from deep-sea thermophiles
253
with typical primer sets - due to mismatches with limited sequence representation in GenBank -
254
it is possible that these assays similarly underestimate abundances in these environments
11
255
(Teske & Sorensen, 2008). Deltaproteobacterial abundance at Needles was approximately 4.48
256
x 105 copies g-1 sample (approximately 26% of the entire bacterial population), though none
257
were detected at Chowder Hill or Dead Dog (data not shown). The abundance of both
258
functional genes for sulfate reduction, dsrA and aprA, was greatest at Needles and lowest at
259
Dead Dog (a pattern similar to that seen in the 16S rRNA gene abundance estimates; Figure 2B).
260
If we assume an average of 1 dsrA gene copy per genome (Klein et al. 2001; Kondo et al. 2004),
261
the proportion of sulfate reducing bacteria in the bacterial populations is only 2.7% in Needles
262
as compared with 28% in Dead Dog and 53% at Chowder Hill.
263
Microbial Diversity
264
454 pyrotag sequencing (bacterial V1-V3 and archaeal V3-V4 of the 16S rRNA gene),
265
rarefaction analyses, and diversity metrics all revealed measureable differences in microbial
266
community composition among the three hydrothermal deposits (Figure 3, Figure 4, Table 3).
267
Via these assessments, Needles hosts the least diverse assemblage of bacteria and archaea,
268
while Chowder Hill and Dead Dog host communities of comparable diversity. Examination of
269
OTUs at 97%, 95% and 92% sequence similarity further reveal differences in microbial
270
community membership among the three sites. Among archaea at the 97% level, only two
271
archaeal OTUs (1% of all archaeal OTUs classified) are shared among the hydrothermal
272
deposits. The sequences classified to these OTUs represent 69%, 48% and 18% of all the library
273
sequences from Needles, Dead Dog and Chowder Hill respectively. One of these OTUs is allied
274
to the ammonium oxidizing archaeal Candidatus Cenarchaeum in the phylum Thaumarchaeota,
275
and accounts for 35% of Needles and less than 5.0% of Dead Dog or Chowder Hill library
276
sequences. The other OTU is allied to a thermophilic sulfur respiring archaeon within the class
12
277
Thermoplasmata. Nearly 40% of the archaeal sequences from Dead Dog were allied to this
278
archaeon. Methanogens allied to Methanocaldococus comprised about 1.0% of the total
279
archaeal sequences from Dead Dog, and were not represented in the libraries from Chowder
280
Hill or Needles. Most of the archaeal diversity at Chowder Hill (80% of sequences) and Dead
281
Dog (50% of sequences) was unclassified. No sequences allied to true sulfate reducing archaeal
282
lineages such as Archaeoglobus fugilis or Aciduliprofundus boonei were recovered. However,
283
the potential diversity of thermophilic sulfate reducing archaea in these samples is likely much
284
greater than suggested here. This may be explained in part by biases underlying DNA
285
extractions, primer binding and sequencing. For example the archaeal sequencing primers used
286
in this study only target about one third (34%, as assayed by Probe Match; Cole et al. 2005) of
287
the Archaeoglous-like sequences contained in the RDP database. Furthermore, the primers may
288
miss members of the dominant Thermoplasmatales as in silico analysis only returns 48%
289
(1715/3558 sequences) of the RDP reported sequences. In total, these archaeal sequencing
290
primers (349F-806R) miss 42% of the total archaeal sequences (67713/117373 sequences) in
291
the RDP database. Similar bias has been reported in other studies in the deep sea and deep
292
subsurface biotopes (e.g. Dhillion et al., 2003, 2005, Teske et al. 2007).
293
Among bacteria at the 97% similarity level, 54 of the bacterial OTUs classified (7.0%)
294
were shared among all hydrothermal deposits, and account for 84%, 80%, 71% of the
295
sequences from Chowder Hill, Needles and Dead Dog respectively (Figure4). One of these OTUs
296
accounted for 44%, 36% and 25% of the sequences from Chowder Hill, Dead Dog and Needles
297
respectively. Aligning representative sequences from this OTU via Blast-n (Altschul et al. 1997)
298
reveals a best match to Thermodesulfovibrio, an anaerobic, thermophilic, sulfate-reducing
13
299
bacterium, from the phylum Nitrospira (81% identity). Given its abundance, we postulate that it
300
likely contributes substantially to the high thermophilic sulfate reduction rates. Furthermore,
301
the
302
Thermodesulfovibrio (Supplemental Table S1). Other dominant groups of bacteria include
303
members of the
304
proteobacteria. Via Blast-n, most of the unclassified sequences matched to partial 16S rRNA
305
gene sequences from hydrothermal vent fluid communities (Nunoura et al. 2010; Sylvan et al.
306
2012). Sequences classified as Deltaproteobacteria comprised 5.5%, 8.2%, or 14%, of the total
307
population of Chowder Hill, Needles, or Dead Dog respectively. While Dead Dog may have had
308
the highest proportion of its amplicons classified as Deltaproteobacteria, sequences related to
309
known sulfate reducers within the Deltaproteobacteria (Desulfobacteraceae, Desulfobulbus
310
rhabdoformis, Desulfovibrio) were only found at Needles and comprised 1.1% of the 16S rRNA
311
gene library. The majority of the sequences classified as Deltaproteobacteria in each of the
312
three sites were from one unclassified Deltaproteobacterial OTU consisting of 4.4, 5.3, and 13%
313
of the sequences from Chowder Hill, Needles and Dead Dog respectively.
314
DISCUSSION
dsrB
gene
library
was
dominated
Gammaproteobacteria,
by
sequences
Bacteroidetes,
phylogenetically
Deltaproteobacteria
allied
and
to
α-
315
Sulfate reduction rates measured in deposits recovered from the Middle Valley vent
316
field reveal the potential for active sulfate reduction within hydrothermal deposits. The
317
magnitude of all measured rates (from 15.7 nmol g-1 day-1 at Dead Dog at 60°C to 2670 nmol g-1
318
day-1 at Needles at 90°C) under experimental conditions were markedly higher than those
319
typically observed in non hydrothermal deep sea sediments (0.1-10 nmol g-1 day-1, converted
320
here for comparison assuming an average sediment density of 2 g cm-3; Elsgaard, Isaksen, et al.
14
321
1994; Weber & Jorgensen, 2002; Joye et al. 2004). These rates are comparable in magnitude to
322
those previously observed in hydrothermally influenced sediments (eg. Guaymas basin or Lake
323
Tanganyika (Weber & Jorgensen, 2002; Kallmeyer & Boetius, 2004; Elsgaard, Isaksen, et al.
324
1994; Elsgaard, Prieur, et al. 1994), even though the availability of organic carbon is markedly
325
higher in these hydrothermal vent sediments, with Guaymas having up to 200-times greater
326
concentrations of organic carbon (Lang et al. 2006; Cruse & Seewald, 2006; Chen et al. 1993).
327
To date, the only other measurement of sulfate reduction from sulfide deposits along the East
328
Pacific Rise exhibited rates comparable to those reported here, (Bonch-Osmolovskaya et al.
329
2011), but it should be noted that their samples were incubated under pure H2 atmosphere.
330
Notably, the maximum rates of sulfate reduction in Middle Valley sulfides occurred at
331
90°C in all three deposits. This is in contrast to measurements of sulfate reduction in
332
hydrothermal sediments, where the greatest rates are often observed between 40-70°C, and
333
more modest rates of sulfate reduction have been reported between 80-91°C (Weber &
334
Jorgensen, 2002; Elsgaard, Prieur, et al. 1994; Elsgaard, Isaksen, et al. 1994). The relatively low
335
or insignificant sulfate reduction rates between 4-80°C suggest Middle Valley deposits harbor a
336
high proportion of hyperthermophilic sulfate-reducing microbes.
337
The significant differences in the rates we observed among deposits (Kruskal-Wallis,
338
p<0.0001) are likely due to differences in biomass and the composition of microbial
339
communities that are influenced by the geochemistry of each deposit. Indeed, microbial
340
biomass (as estimated by 16S rRNA genes) directly correlates to rates of activity and is likely
341
one of the strongest factors affecting the observed rates of sulfate reduction (Pearson
342
correlation coefficient r=0.879, p<0.0005). Needles had both the highest observed rates as well
15
343
as the highest cell density (Figure 2 and 3). Of all the deposits sampled, Needles had the lowest
344
venting fluid temperature (123°C) resulting in the largest zone of microbial habitability.
345
Consistent with this, Needles also had the greatest abundance of dsrA and aprA genes per
346
gram, suggesting a larger potential sulfate reducing community. Here, Deltaproteobacteria
347
allied to Desulfovibrio, Desulfobulbus, Desulfobacteria, and Desulfuromonas account for 25.7%
348
of the bacterial community. These clades of Deltaproteobacteria were not observed at
349
Chowder Hill or Dead Dog by either qPCR enumeration or pyrosequencing. Cultured
350
representatives from some of these Deltaproteobacterial clades (Desulfovibrio vulgaris and
351
Desulfovibrio desulfuricans) have been shown to reduce sulfate at high rates (ranging from 10-
352
1340 nmol min-1 mg-1 protein) with varying electron donors (Fitz & Cypionka, 1991; Cypionka &
353
Konstanz, 1989).
354
Thermodesulfovibrio-like organisms dominated the bacterial communities within each
355
hydrothermal deposit (35-44%; Fig. 4). Thermodesulfovibrio sp. are considered obligately
356
anaerobic, thermophilic bacteria that can reduce sulfate and other sulfur compounds (Garrity &
357
Holt, 2001). In pure cultures, members of this genus are able to link growth with hydrogen and
358
a limited range of organic carbon molecules (formate, pyruvate and lactate), maintaining
359
optimal growth between 55-70°C (Sekiguchi et al., 2008). Needles has a greater proportion of
360
sequences (from pyrosequencing) related to Thermodesulfovibrio-like species than the other
361
two deposits. The combination of many sequences related to a thermophilic sulfate reducing
362
bacteria and higher rates of sulfate reduction at elevated temperatures and the, suggests that
363
Thermodesulfovibrio-like organisms may play a role in sulfate reduction in warmer
364
environments. However, constraining the relative proportion of sulfate reduction by
16
365
Thermodesulfovibrio-like organisms in these mixed communities was beyond the scope of this
366
study.
367
It is unclear why Chowder Hill and Dead Dog exhibit large differences in rates of sulfate
368
reduction despite other similarities in geochemistry and biomass. One plausible explanation
369
might be that different types of biological interactions (e.g. syntrophy or competition) occur
370
due to differences in the composition and distribution of microbial communities within the
371
mineral matrix of each deposit. Slight differences in community composition, like Dead Dog
372
having a higher representation of sequences related to sulfur respiring (Thermoplasmata) and
373
methanogenic (Methanocaldococus-like) archaea than Chowder Hill, may lead to biological
374
interactions that have different implications for rates of sulfate reduction in each deposit. Also,
375
substrate competition for H2 or consumption of locally produced DOC (Oremland & Polcin,
376
1982; Lovley & Phillips, 1987) may be more prevalent in one deposit over another. Future
377
experiments should aim to better resolve how specific interaction between populations, for
378
example, syntrophy or competition for a common substrate, may influence sulfate reduction.
379
The relevance of heterotrophic sulfate reduction on hydrothermal vent biomass production
380
and biogeochemistry
381
Heterotrophic sulfate reduction is likely a prominent metabolic mode within Middle
382
Valley sulfides and sediments, and the sulfate reduction rate data herein (which solely measure
383
hetrotrophic sulfate reduction) support that supposition. Sedimented vent fields typically
384
contain allochthonous organic carbon that could readily support heterotrophy. Indeed, at
385
Middle Valley, bottom waters contain 3.5 mg DOC/L (about 7 fold higher than the overlying
386
surface seawater), while porewater concentrations range from 0.1 – 84.0 mg DOC/L at
17
387
sediment depths to 200 mbsf (Ran & Simoneit, 1994). Based on data from culture studies of
388
Desulfovibrio strains, including the H+/H2 ratio of 1.0 for Desulfovibrio vulgarius Marburg (Fitz &
389
Cypionka, 1991), the P/2e- ratio (number of ATPs produced for every 2 electrons transferred to
390
an electron acceptor) of 1/3 for Desulfovibrio gigas (Barton et al. 1983), and the assumption
391
that 10% of ATP production supports growth (20 mmol ATP per gram biomass), our estimates
392
suggest that - at our maximum empirically measured rates - heterotrophic sulfate reduction
393
could support 140 g biomass yr-1 (~1.5 x 1014 cells) at Chowder Hill (volume = 109,900 cm3), 16 g
394
biomass yr-1 (~1.7 x 1013 cells) at Needles (volume = 5495 cm3), and 2.1 g biomass yr-1 (~2.2 x
395
1012 cells) at Dead Dog (volume=12560 cm3). While these values may be small in comparison to
396
global estimates of chemoautotrophic biomass production on the global ridge system (1010 -
397
1013 g of biomass yr-1 ;McCollom & Shock 1997; Bach & Edwards 2003), the sulfide produced by
398
these heterotrophic sulfate reducers could represent up to 3% of the H 2S flux from Middle
399
Valley deposits (given previously published vent fluid flow rates from the Main Endeavor field,
400
(Wankel et al. 2011)). Additional rate measurements that represent the diversity of physico-
401
chemical conditions found within deposits or ridge systems are necessary to better constrain
402
the contribution of heterotrophic sulfate reducers to global vent biomass and geochemistry.
403
Hydrothermal vents are dynamic environments where carbon and sulfur cycling are
404
intimately linked. Both autotrophic and heterotrophic sulfate reducing microbes have been
405
isolated from vents, and the data shown here are among the first to constrain the potential for
406
heterotorophic sulfate reduction at vents (in particular those with higher organic carbon loads),
407
as well as the relationship between sulfate reduction rates, temperature, microbial biomass
408
and community density and composition. These data, as well as the vent field estimates of
18
409
sulfate reduction, , underscore the relevance of sulfate reduction in hydrothermal ecosystems
410
and further indicate the need for continued studies of sulfur cycling along ridge systems.
411
412
ACKNOWLEDGMENTS:
413
We are grateful for the expert assistance of the R/V Atlantis crews and the pilots and
414
team of the DSV Alvin for enabling the collections of hydrothermal deposits used in our
415
experiments. We also thank Steve Sansone, Dr. Joseph Ring, Ms. Julie Hanlon, Dr. Kathleen
416
Scott, Dr. Vladimir Samarkin, Dr. David Johnston, and Dr. Jan Amend for providing assistance
417
with various technical aspects of the experiments. We are also very thankful for the
418
constructive feedback from the reviewers. Financial support for this research was provided by
419
the National Science Foundation (NSF OCE-0838107 and NSF OCE-1061934 to P.R. Girguis), and
420
the National Aeronautic and Space Administration (NASA-ASTEP NNX09AB78G to C. Scholin and
421
P. R. Girguis and NASA-ASTEP NNX07AV51G to A. Knoll and P. R.Girguis).
422
Supplementary information is available at ISME’s website
423
424
REFERENCES
425
426
427
Alazard D, Dukan S, Urios A, Verhe F, Bouabida N, Morel F, et al. (2003). Desulfovibrio
hydrothermalis sp nov., a novel sulfate-reducing bacterium isolated from hydrothermal vents.
Int J Syst Evol Microbiol 53:173–178.
428
429
430
Altschul SF, Madden TL, Schäffer a a, Zhang J, Zhang Z, Miller W, et al. (1997). Gapped BLAST
and PSI-BLAST: a new generation of protein database search programs. Nucleic Acids Res
25:3389–3402.
431
432
433
Amend JP, Shock EL. (2001). Energetics of overall metabolic reactions of thermophilic and
hyperthermophilic Archaea and Bacteria. Fems Microbiology Reviews 25:175–243. <Go to
ISI>://000167599700002.
19
434
435
436
Audiffrin C, Cayol JL, Joulian C, Casalot L, Thomas P, Garcia JL, et al. (2003). Desulfonauticus
submarinus gen. nov., sp nov., a novel sulfate-reducing bacterium isolated from a deep-sea
hydrothermal vent. Int J Syst Evol Microbiol 53:1585–1590.
437
438
439
Bach Wolfgang, Edwards Katrina J. (2003). Iron and sulfide oxidation within the basaltic ocean
crust : Implications for chemolithoautotrophic microbial biomass production. Geochimica et
Cosmochimica 67:3871–3887.
440
441
442
Barton LL, Legall J, Odom JM, Peck HD. (1983). Energy coupling to nitrite sulfate-reducing
bacterium Energy Coupling to Nitrite Respiration in the Sulfate- Reducing Bacterium
Desulfovibrio gigas. Society.
443
444
445
Blochl E, Rachel R, Burggraf S, Hafenbradl D, Jannasch HW, Stetter KO. (1997). Pyrolobus
fumarii, gen. and sp. nov., represents a novel group of archaea, extending the upper
temperature limit for life to 113 degrees C. Extremophiles 1:14–21.
446
447
448
449
Bonch-Osmolovskaya E a, Perevalova A a, Kolganova TV, Rusanov II, Jeanthon Christian,
Pimenov NV. (2011). Activity and distribution of thermophilic prokaryotes in hydrothermal
fluid, sulfidic structures, and sheaths of alvinellids (East Pacific Rise, 13°N). Appl. Environ.
Microbiol. 77:2803–2806.
450
451
452
Bostick BC, Fendorf S, Helz GR. (2003). Differential adsorption of molybdate and
tetrathiomolybdate on pyrite (FeS2). Environmental science & technology 37:285–91.
http://www.ncbi.nlm.nih.gov/pubmed/12564899.
453
454
455
456
Butterfield David A, McDuff RE, Mottl Michael J, Lilley Marvin D, Lupton JE, Massoth GJ. (1994).
Gradients in the composition of hydrothermal fluids from the Endeavour segment vent field:
Phase separation and brine loss. J. Geophys. Res. 99:9561–9583.
http://dx.doi.org/10.1029/93JB03132.
457
458
Butterfield David A, Mcduff RE, Franklin J, Wheat CG. (1994). Geochemistry of hydrothermal
vent fluids from Middle Valley, Juan de Fuca ridge. Proc ODP Sci Res 139:395–410.
459
460
Campbell BJ, Engel AS, Porter ML, Takai Ken. (2006). The versatile epsilon-proteobacteria: key
players in sulphidic habitats. Nature Rev Microbiol 4:458–68.
461
462
463
Canfield DE. (1989). Sulfate Reduction and oxic respiration in marine- sediments- Implications
for organic-carbon preservation in euxinic envrionments. Deep Sea Research Part A
Oceanographic Research Papers 36:121–138.
464
465
466
Chen RF, Bada JL, Suzuk Y. (1993). The relationship between dissolved organic carbon ( DOC )
and fluorescence in anoxic marine porewaters: Implications for estimating benthic DOC fluxes.
Geochimica Et Cosmochimica Acta 57:2149–2153.
20
467
468
469
Christophersen CT, Morrison M, Conlon MA. (2011). Overestimation of the abundance of
sulfate-reducing bacteria in human feces by quantitative PCR targeting the Desulfovibrio 16S
rRNA gene. Appl. Environ. Microbiol. 77:3544–3546.
470
471
472
473
474
Cole JR, Chai B, Farris RJ, Wang Q, Kulam S a, McGarrell DM, et al. (2005). The Ribosomal
Database Project (RDP-II): sequences and tools for high-throughput rRNA analysis. Nucleic acids
research 33:D294–6.
http://www.pubmedcentral.nih.gov/articlerender.fcgi?artid=539992&tool=pmcentrez&rendert
ype=abstract (Accessed November 13, 2012).
475
476
477
Cruse AM, Seewald JS. (2006). Geochemistry of low-molecular weight hydrocarbons in
hydrothermal fluids from Middle Valley, northern Juan de Fuca Ridge. Geochimica Et
Cosmochimica Acta 70:2073–2092.
478
479
Cypionka H, Konstanz U. (1989). Characterization of sulfate transport in Desulfovibrio
desulfuricans. Arch. Microbiol. 152:237–243.
480
481
482
Dowd S E, Sun Y, Secor PR, Rhoads DD, Wolcott BM, James GA, et al. (2008). Survey of bacterial
diversity in chronic wounds using Pyrosequencing, DGGE, and full ribosome shotgun
sequencing. BMC Microbiology 8.
483
484
Elsgaard L, Guezennec J, Benbouzidrollet N, Prieur D. (1995). Mesophilic sulfate-reducing
bacteria from 3 deep-sea hydrothermal vent sites. Oceanologica Acta 18:95–104.
485
486
487
Elsgaard L, Isaksen MF, Jorgensen B B, Alayse AM, Jannasch HW. (1994). Microbial sulfate
reduction in deep-sea sediments at the Guaymas Basin hydrothermal vent area: Influence of
temperature and substrates. Geochimica Et Cosmochimica Acta 58:3335–3343.
488
489
Elsgaard L, Prieur D, Mukwaya GM, Jorgensen B B. (1994). Thermophilic sulfate reduction in
hydrothermal sediment of lake tanganyika, East Africa. Appl. Environ. Microbiol. 60:1473–1480.
490
491
Fitz R, Cypionka H. (1991). Generation of a proton gradient in Desulfovibrio vulgaris. Arch.
Microbiol. 155:444–448.
492
493
494
Fossing H, Jorgensen Bo Barker. (1989). Chromium Reduction Method of bacterial sulfate
reduction in sediments: Measurement reduction of a single-step chromium method Evaluation.
Biogeochemistry 8:205–222.
495
496
497
Frias-lopez J, Zerkle AL, Bonheyo GT, Fouke BW. (2002). Partitioning of Bacterial Communities
between Seawater and Healthy , Black Band Diseased , and Dead Coral Surfaces. Appl. Environ.
Microbiol. 68:2214–2228.
498
499
Garrity G.M., Holt JG. (2001). The road map to the Manual. In:Bergys Manual of Systematic
Bacteriology, Castenholz, RW & Garrity, G.M, eds (ed)., Springer: New York, pp. 119–166.
21
500
501
502
Hoek J, Banta A, Hubler F, Reysenbach AL. (2003). Microbial diversity of a sulphide spire located
in the Edmond deep-sea hydrothermal vent field on the Central Indian Ridge. Geobiology
1:119–127.
503
504
505
Houghton JL, Seyfried WE, Banta AB, Reysenbach AL. (2007). Continuous enrichment culturing
of thermophiles under sulfate and nitrate-reducing conditions and at deep-sea hydrostatic
pressures. Extremophiles 11:371–382.
506
507
Huber JA, Mark Welch D, Morrison HG, Huse SM, Neal PR, Butterfield D A, et al. (2007).
Microbial population structures in the deep marine biosphere. Science 318:97–100.
508
509
Jannasch HW, Mottl M J. (1985). Geomicrobiology of Deep-sea Hydrothermal Vents. Science
229:717–725.
510
511
512
Jannasch HW, Wirsen CO, Molyneaux SJ, Langworthy TA. (1988). Extremely thermophilic
fermentative archaebacteria of the genus Desulfurococcus from deep-sea hydrothermal vents.
Appl. Environ. Microbiol. 54:1203–1209.
513
514
Johnson KS, Beehler CL, Sakamoto-Arnold C., Childress J. (1986). In situ measurements of
chemical distributions in a deep-sea hydrothermal vent field. Science 231:1139–1141.
515
516
Jorgensen B B, Isaksen MF, Jannasch HW. (1992). Bacterial sulfate reduction above 100degrees-C in deep-sea hydrothermal vent sediments. Science 258:1756–1757.
517
518
519
Joye SB, Boetius A, Orcutt BN, Montoya JP, Schulz HN, Erickson MJ, et al. (2004). The anaerobic
oxidation of methane and sulfate reduction in sediments from Gulf of Mexico cold seeps.
Chemical Geology 205:219–238.
520
521
522
523
Kallmeyer J, Boetius Antje. (2004). Effects of Temperature and Pressure on Sulfate Reduction
and Anaerobic Oxidation of Methane in Hydrothermal Sediments of Guaymas Basin Effects of
Temperature and Pressure on Sulfate Reduction and Anaerobic Oxidation of Methane in
Hydrothermal Sediments of. Appl. Environ. Microbiol. 70:1231–1233.
524
525
Klappenbach J a, Saxman PR, Cole JR, Schmidt T M. (2001). rrndb: the Ribosomal RNA Operon
Copy Number Database. Nucleic Acids Res 29:181–184.
526
527
528
529
Klein M, Friedrich M, Roger Andrew J, Hugenholtz P, Fishbain S, Abicht H, et al. (2001). Multiple
Lateral Transfers of Dissimilatory Sulfite Reductase Genes between Major Lineages of SulfateReducing Prokaryotes. Appl. Environ. Microbiol. 183:6028–6035. <Go to
ISI>://000171267100028.
530
531
Kondo R, Nedwell David B., Purdy KJ, Silva SQ. (2004). Detection and Enumeration of SulphateReducing Bacteria in Estuarine Sediments by Competitive PCR. Geomicrobiol J 21:145–157.
22
532
533
534
Lang SQ, Butterfield D A, Lilley M D, Johnson HP, Hedges JI. (2006). Dissolved organic carbon in
ridge-axis and ridge-flank hydrothermal systems. Geochimica Et Cosmochimica Acta 70:3830–
3842.
535
536
Lee ZM-P, Bussema C, Schmidt Thomas M. (2009). rrnDB: documenting the number of rRNA and
tRNA genes in bacteria and archaea. Nucleic Acids Res 37:489–493.
537
538
539
Lovley D R, Phillips EJ. (1987). Competitive mechanisms for inhibition of sulfate reduction and
methane production in the zone of ferric iron reduction in sediments. Appl. Environ. Microbiol.
53:2636–2641.
540
541
542
Manefield M, Whiteley AS, Griffiths RI, Bailey MJ. (2002). RNA Stable Isotope Probing , a Novel
Means of Linking Microbial Community Function to Phylogeny. Appl. Environ. Microbiol.
68:5367–5373.
543
544
545
McCollom TM, Shock EL. (1997). Geochemical constraints on chemolithoautotrophic
metabolism by microorganisms in seafloor hydrothermal systems. Geochimica Et Cosmochimica
Acta 61:4375–4391.
546
547
Muyzer G, Stams AJM. (2008). The ecology and biotechnology of sulphate-reducing bacteria.
Nature Rev Microbiol 6:441–544.
548
549
550
Nakagawa Satoshi, Takai K, Inagaki F, Hirayama H, Nunoura T, Horikoshi K, et al. (2005).
Distribution, phylogenetic diversity and physiological characteristics of epsilon-Proteobacteria
in a deep-sea hydrothermal field. Environ. Microbiol. 7:1619–1632.
551
552
Nakagawa Satoshi, Takai Ken. (2008). Deep-sea vent chemoautotrophs: diversity, biochemistry
and ecological significance. FEMS Microbiol Ecol 65:1–14.
553
554
555
Nakagawa T, Nakagawa S, Inagaki F, Takai K, Horikoshi K. (2004). Phylogenetic diversity of
sulfate-reducing prokaryotes in active deep-sea hydrothermal vent chimney structures. Fems
Microbiol Let 232:145–152.
556
557
558
Nercessian O, Bienvenu N, Moreira D, Prieur D, Jeanthon C. (2005). Diversity of functional genes
of methanogens, methanotrophs and sulfate reducers in deep-sea hydrothermal environments.
Environ. Microbiol. 7:118–132.
559
560
Newport PJ, Nedwell D B. (1988). The mechanisms of inhibition of Desulfovibrio and
Desulfotomaculum species by selenate and molybdate. J Appl Microbiol 65:419–423.
561
562
563
564
Nunoura Takuro, Oida H, Nakaseama M, Kosaka A, Ohkubo SB, Kikuchi T, et al. (2010). Archaeal
diversity and distribution along thermal and geochemical gradients in hydrothermal sediments
at the Yonaguni Knoll IV hydrothermal field in the Southern Okinawa trough. Appl. Environ.
Microbiol. 76:1198–1211.
23
565
566
567
Oakley BB, Carbonero F, Dowd Scot E, Hawkins RJ, Purdy KJ. (2011). Contrasting patterns of
niche partitioning between two anaerobic terminal oxidizers of organic matter. ISME J 6:905–
914.
568
569
Oremland RS, Polcin S. (1982). Methanogenesis and sulfate reduction: competitive and
noncompetitive substrates in estuarine sediments. Appl. Environ. Microbiol. 44:1270–1276.
570
571
Ran B, Simoneit BRT. (1994). Dissolved organic carbon in interstitial waters from sediments of
Middle Valley, Leg 139 1. Proc ODP Sci Res 139:441–446.
572
573
Saleh AM, Macpherson R, Miller JDA. (1964). The Effect of Inhibitors on Sulphate Reducing
Bacteria: a Compilation. J Appl Microbiol 27:281–293.
574
575
Santelli CM, Orcutt BN, Banning E, Bach W, Moyer CL, Sogin ML, et al. (2008). Abundance and
diversity of microbial life in ocean crust. Nature 453:653–657.
576
577
578
Schloss PD, Westcott SL, Ryabin T, Hall JRA, Hartmann M, Hollister EB, et al. (2009). Introducing
mothur: open-source, platform-independent, community-supported software for describing
and comparing microbial communities. Appl. Environ. Microbiol. 75:7537–7541.
579
580
581
Schrenk MO, Kelley DS, Delaney JR, Baross JA. (2003). Incidence and Diversity of
Microorganisms within the Walls of an Active Deep-Sea Sulfide Chimney. Appl. Environ.
Microbiol. 69:3580–3592.
582
583
584
585
586
Sekiguchi Y, Muramatsu M, Imachi H, Narihiro T, Ohashi A, Harada H, et al. (2008).
Thermodesulfovibrio aggregans sp. nov. and Thermodesulfovibrio thiophilus sp. nov.,
anaerobic, thermophilic, sulfate-reducing bacteria isolated from thermophilic methanogenic
sludge, and emended description of the genus Thermodesulfovibrio. Int J Syst Evol Microbiol
58:2541–2548.
587
588
589
Stults J R, Snoeyenbos-West O, Methe B, Lovley D R, Chandler D P. (2001). Application of the 5 `
fluorogenic exonuclease assay (TaqMan) for quantitative ribosomal DNA and rRNA analysis in
sediments. Applied and Environmental Microbiology 67:2781–2789.
590
591
592
593
594
Stults Jennie R, Snoeyenbos-west O, Methe Barbara, Lovley Derek R, Chandler Darrell P. (2001).
Application of the 5 ′ Fluorogenic Exonuclease Assay ( TaqMan ) for Quantitative Ribosomal
DNA and rRNA Analysis in Sediments Application of the 5Ј Fluorogenic Exonuclease Assay (
TaqMan ) for Quantitative Ribosomal DNA and rRNA Analysis in Sediments. Appl. Environ.
Microbiol. 67:2781–2789.
595
596
Suzuki MT, Beja O, Taylor LT, DeLong EF. (2001). Phylogenetic analysis of ribosomal RNA
operons from uncultivated coastal marine bacterioplankton. Environ. Microbiol. 3:323–331.
24
597
598
Sylvan JB, Toner BM, Edwards Katrina J. (2012). Life and Death of Deep-Sea Vents : Bacterial
Diversity and Ecosystem Succession on Inactive Hydrothermal Sulfides. mBio 3:e00279–11.
599
600
601
Takai KEN, Horikoshi Koki. (2000). Rapid Detection and Quantification of Members of the
Archaeal Community by Quantitative PCR Using Fluorogenic Probes. Appl. Environ. Microbiol.
66:5066–5072.
602
603
Teske A, Sorensen KB. (2008). Uncultured archaea in deep marine subsurface sediments: have
we caught them all? ISME J 2:3–18.
604
605
Wagner M, Roger A J, Flax JL, Brusseau GA, Stahl D A. (1998). Phylogeny of dissimilatory sulfite
reductases supports an early origin of sulfate respiration. J Bact 180:2975–2982.
606
607
608
Wankel SD, Germanovich LN, Lilley Marvin D., Genc G, DiPerna CJ, Bradley AS, et al. (2011).
Influence of subsurface biosphere on geochemical fluxes from diffuse hydrothermal fluids.
Nature Geo 4:461–468.
609
610
611
Weber A, Jorgensen B B. (2002). Bacterial sulfate reduction in hydrothermal sediments of the
Guaymas Basin, Gulf of California, Mexico. Deep-Sea Research Part I-Oceanographic Research
Papers 49:827–841.
612
613
614
Xu N, Christodoulatos C, Braida W. (2006). Adsorption of molybdate and tetrathiomolybdate
onto pyrite and goethite: effect of pH and competitive anions. Chemosphere 62:1726–35.
http://www.ncbi.nlm.nih.gov/pubmed/16084558 (Accessed November 21, 2012).
615
616
617
(Christophersen et al. 2011; Frias-lopez et al. 2002; Oakley et al. 2011; M Wagner et al. 1998;
Manefield et al. 2002; Suzuki et al. 2001; Ken E N Takai & Koki Horikoshi 2000)
618
619
25
Download